You are on page 1of 8

American Journal of Hematology 78:232239 (2005)

Current Views in HTLV-I-Associated Adult T-Cell


Leukemia/Lymphoma
Christophe Nicot
Department of Microbiology, Immunology and Molecular Genetics, University of Kansas Medical Center, Kansas City, Kansas

Epidemiological studies have demonstrated that the relative percentage of malignant lymphoid proliferations varies widely according to geographical location and ethnic populations. HTLV-I is the etiological agent of adult T-cell leukemia/lymphoma (ATLL) and is also
associated with cutaneous T-cell lymphoma (CTCL). However, a definite role of HTLV-I in
mycosis fungoides (MF) and/or Sezary syndrome (SS) remains controversial. While most
HTLV-I-infected individuals remain asymptomatic carriers, 15% will develop ATLL, an invariably fatal expansion of virus-infected CD4+ T cells. This low incidence and the long latency
period preceding occurrence of the disease suggest that additional factors are involved in
development of ATLL. In this review, diagnosis, clinical features, and molecular pathogenesis of HTLV-I are discussed. Am. J. Hematol. 78:232239, 2005. 2005 Wiley-Liss, Inc.

DISCOVERY AND EPIDEMIOLOGY

In 1977, epidemiological studies revealed the presence of unusual clusters of adult T-cell leukemia/
lymphoma (ATLL) in some areas of Japan, suggesting
a transmissible agent may be involved in the disease [1].
The first description of HTLV-I came after the discovery of the human T-cell growth factor (interleukin-2;
IL-2) [2], allowing long-term in vitro culture of T cells
and establishment of T-cell lines from a patient with a
cutaneous T-cell lymphoma (CTCL) [3]. Soon after,
this virus was identified as the etiological agent of
ATLL [4], and the term HTLV-I was adopted.
The relative percentage of malignant lymphoid
proliferations varies widely according to geographical
location, probably reflecting exposure to different
etiological factors, including viruses. Peripheral T-cell
lymphoma (PTCL) is relatively uncommon in Caucasian populations. For patients with non-Hodgkin lymphoma (NHL), the proportion of PTCL is only around
10% in Western countries [5] while the incidence of
PTCL is as high as 70% in southwestern Japan, where
HTLV-I infection is endemic. A recent survey of
lymphoid malignancies in Japan showed that 50%
were B-cell lymphoma and 42% were T/natural killer
(NK)-cell lymphoma, among which 58% were HTLV-I
infected while only 4% were Hodgkin lymphoma
(HL) [6].
2005 Wiley-Liss, Inc.

All routes of HTLV-1 virus transmission require


close contact with infected T-lymphocytes. Motherto-child transmission is associated with prolonged
breast-feeding in the postnatal period [710] and has
been associated with an increased risk of developing
ATLL. HTLV-1 can be sexually transmitted with a
higher transmission efficiency from male to female
than from female to male [11,12]. The intravenous
route of infection, mainly by blood transfusion,
appears to be the most efficient mode for HTLV-1
transmission [13]. The intravenous route of contamination is associated with a higher risk of developing tropical spastic paraparesis/HTLV-1 associated
Contract grant sponsor: National Institutes of Health; Contract
grant number: RR016443 (NIH) from the COBRE Program of the
National Center for Research Resources; Contract grant sponsor:
Lied Basic Science Research Grant of the University of Kansas
Medical Center Research Institute.
*Correspondence to: C. Nicot, University of Kansas Medical
Center, Department of Microbiology, Immunology and
Molecular Genetics, 3025 Wahl Hall West, 3901 Rainbow
Boulevard, Kansas City, KS 66160. E-mail: cnicot@kumc.edu
Received for publication 28 May 2004; Accepted 29 September
2004
Published online in Wiley InterScience (www.interscience.wiley.com).
DOI: 10.1002/ajh.20307

Concise Review: Current Views in HTLV-I-Associated Adult T-Cell Leukemia/Lymphoma

myelopathy (TSP/HAM). Transmission of HTLV-I


by blood transfusion occurs with transfusion of cellular blood products (whole blood, red blood cells,
and platelets) but not with the plasma fraction or
plasma derivatives from HTLV-I-infected blood. Seroconversion rates of 4463% have been reported in
recipients of HTLV-I-infected cellular components in
HTLV-I endemic areas.
In 1988, the Food and Drug Administration (FDA)
recommended all blood donation centers screen the
U.S. blood supply for HTLV-I. In 2001, data relating
to all blood donations to the American Red Cross
indicated the rates per 100,000 were 9.7 for HIV and
9.6 for HTLV. Incidences of new infection among
donors were 1.554 and 0.239 for HIV and HTLV,
respectively. Current estimations indicate 2030 million people worldwide are infected with HTLV-I.
Endemic areas are mainly found in Japan, Africa,
South America, Caribbean basin, Southern parts of
North America, and Eastern Europe. While most
HTLV-I-infected individuals remain asymptomatic
carriers, 15% (lifelong risk) will develop ATLL, an
invariably fatal expansion of virus-infected CD4+
T cells. The human T-cell lymphotropic virus type II
(HTLV-II) resembles HTLV-I in provirus organization and exerts less lymphoproliferative effects on the
hosts infected cells. To date, a clear association of
HTLV-II with a specific human disease has not been
established.
DIAGNOSIS CRITERIA

The diagnosis of ATL is usually made on morphological analysis. Cytological examination may reveal
infiltration by cerebriform or flower cells (activated
lymphocytes with convoluted nuclei and basophilic
cytoplasm), indicators of acute or lymphoma type
ATL. This must be confirmed by clonal integration of
HTLV-I provirus in the host genome. The predominant
immunological phenotype of neoplastic cells is helper
T-cell, CD3+, CD4+, L-selectin+, CD25+, CD45RA+,
HLA-DR+, CD29 , and CD45RO in peripheral
blood, or CD3+, CD4+, L-selectin+, CD29+,
CD45RO+, HLA-DR+, and CD45RA in the skin
and lymph nodes [14,15]. Phenotypic as well as morphological heterogeneity of ATL cells and heterogeneity of CD45R isoform expression on ATL cells can be
found in different organs. The CD7 and CD8 antigens
are usually absent. Factors signifying a poor prognosis
include high serum thymidine kinase levels [16,17], high
serum soluble interleukin-2 receptor levels [18], high
serum b2-microglobulin levels [19], high expression
of the Ki67 antigen, and high serum parathyroid
hormone-related protein levels. The serum neuronspecific enolase (NSE) value positively correlated with

233

serum thymidine kinase activity and serum soluble


interleukin-2 receptor levels. Because ATL cells produce significantly more NSE than other NHL cells,
serum NSE may serve as a marker of disease aggressiveness as well as a prognostic factor for ATL [20].
Whether HTLV-1-associated malignancies are referred
to as leukemias or lymphomas depends on specific
criteria found in the peripheral blood. Acute ATL is
characterized by a massive infiltration of the peripheral
blood by ATL cells, while the ATL lymphoma is characterized by the presence of less than 1% of leukemic
cells on a blood smear and major involvement of lymphoid organs.
The T-cell receptor expressed on leukemic cells is
usually a heterodimer of a and b chains. There are no
reports showing that particular variable segments of
the b chain genes are preferentially expressed in the
leukemic cells of ATLL patients, indicating leukemic
cells are not derived from particular antigen-specific
T-cell clones. According to the Revised European
American Lymphoma (REAL) classification of nonHodgkin lymphomas (NHL), with emphasis on
immunophenotypic analysis along with clinical features, ATLL belongs to the category of peripheral
T-cell lymphoma (PTCL) and natural killer-cell neoplasms. Shirono et al. proposed that ATL is characterized as acute when more than 18% of PBMCs are
found to be Ki-67 antigen-positive [21].
No specific karyotypic abnormalities have been
associated with the development of ATLL, but cytogenetic analyses of leukemic cells revealed multiple
abnormalities such as trisomy 3, 7, and 21, involvement of chromosomes 6 and 14, and loss of chromosome Y [2225]. Recently, loss of heterozygosity on
chromosome 6q [region (6q1521)] was reported in
approximately 50% of acute/lymphoma ATL, suggesting the presence in this location of a putative
tumor-suppressor gene involved in ATLL pathogenesis [26]. The cytogenetic abnormalities found in
ATLL are more frequent in the acute and lymphoma
types than in the chronic or the smoldering types.
Whether or not these genetic alterations are the
cause or the result of ATL is unknown.
HTLV-I-Associated Cutaneous T-Cell Lymphoma
(CTCL)

ATLL prevalence is often underestimated due to the


severity and the rapid evolution of this disease, and the
confusion of ATLL with Sezary syndrome (SS), mycosis fungoides (MF), or other types of T-cell NHL.
There are marked similarities between the clinical and
histopathological features of the lymphoma type of
ATLL and other types of cutaneous T-cell lymphoma
(CTCL) including SS, MF, and CD30 anaplastic large

234

Concise Review: Nicot

cell CTCL. Importantly, liver invasion by ATLL cells


and impaired hepatic functions are frequent in HTLVI-infected individuals as opposed to NHL that are not
associated with HTLV-I infection. Most patients with
CTCL are negative for antibodies to the structural
proteins of HTLV-I, and thus a causative role for this
virus is usually dismissed. However, numerous studies
have found HTLV-I-related proviral tax sequence or
deleted proviruses in tumor samples from CTCL
patients [2734]. Recently a study on the northeastern
coast of Brazil found HTLV-I integrated proviral
sequences by southern blot in several mycosis fungoides-like cutaneous lymphomas, and CD30+ largecell anaplastic lymphomas [35]. However, other studies
could not identify any HTLV-I sequences in MF or SS
patients, and therefore a causative role for HTLV-I
remains controversial [3640].
It is well accepted that HTLV-I is associated with a
cutaneous-lymphoma type of ATL, in which detection
of monoclonal integration of HTLV-I and T-cell
monoclonality can be detected in the skin, but usually
not in the peripheral lymphocytes of ATL patients
[4144]. Various cutaneous lesions have been described
in ATL lymphoma, including papules, nodules, erythroderma, plaques, tumors, and ulcerative lesions. The
frequent expression of the chemokine (C-C motif)
receptor 4 (CCR4) on the surface of tumor cells may
in part explain skin involvement [45,46]. Other factors
such as expression of cutaneous lymphocyte antigen
(CLA) [47,48], expression of lymphocyte chemoattractant chemokine, stromal cell-derived factor/pre-B-cell
growth-stimulating factor (SDF-1/PBSF) [45], as well
as inflammatory responses may lead to chemotaxis of
infected lymphocytes [49] and contribute to skin infiltration by ATL cells. Virus-infected cells can remain in
the skin for several months or years before their dissemination to the peripheral blood and organs.
Clinical Features of HTLV-I-Associated ATL
and ATL Lymphoma

In addition to ATLL, HTLV-I is also the etiological


agent of an inflammatory neurodegenerative disorder
called tropical spastic paraparesis/HTLV-I-associated
myelopathy (TSP/HAM), as well as HTLV-I-associated arthropathy (HAAP), HTLV-I-associated uveitis (HAU), infective dermatitis, and polymyositis. The
role of HTLV-I in TSP/HAM pathogenesis has
recently been reviewed [5052].
ATLL has a broad clinical spectrum divided into four
clinically distinct entities (acute, chronic, smoldering,
and lymphoma) that differ in their presentation, progression and response to treatment [53].
Acute ATLL is characterized by fever, cough, lymphoadenopathy, skin lesions, hepatosplenomegaly,

marked leukocytosis, and hypercalcemia frequently


associated with lytic bone lesions and generalized
bone resorbtion. About 70% of ATL patients develop
high serum calcium levels during the clinical course of
the disease, particularly during the aggressive stage
[54,55]. High serum levels of lactate dehydrogenase, a
soluble form of the interleukin-2 receptor a chain,
and atypical lymphocytes with characteristic convoluted or lobulated nuclei and basophilic cytoplasm,
are also characteristic of the acute type of ATL. The
mean survival time is 6 months with a poor response
to chemo- or radiotherapy.
Chronic-type ATL is characterized by milder clinical symptoms and signs and a longer clinical course.
Serum calcium levels are normal and there is no
organ involvement other than lymphadenopathy,
cutaneous or pulmonary lesions, and hepatosplenomegaly. Chronic lymphocytosis, with more than 10%
circulating leukemia cells and a tendency to be less
cytologically atypical than in the acute type, is also an
indicator of chronic type ATL. The mean survival
time is 24 months.
The smoldering type is characterized by few leukemia
cells in the peripheral blood (less than 5%) and may
present with skin lesions such as papules, nodules, and
erythema. Lymph node enlargement and splenomegaly
are minimal, and serum lactate dehydrogenase levels
are either slightly elevated or normal; hypercalcemia is
usually not detected. Survival is quite long. Both
chronic and smoldering types can progress into an
acute form of leukemia or lymphoma following a progressive aggravation of the clinical picture. Progression
appears to be linked with an increase from low to high
proviral loads, possibly through exposure to chronic
antigen stimulation.
Lymphoma type ATLL is predominantly characterized by lymph node enlargement without manifestations of leukemia. Peripheral blood lacks absolute
lymphocytosis, but sporadic circulating leukemia cells
may be seen (<1%); hypercalcemia is absent. Mean
survival time for this group of patients is 10 months,
and response to chemotherapy is generally poor. The
4-year survival rates are 5.0% for the acute type,
5.7% for the lymphoma type, 26.9% for the chronic
type, and 62.8% for the smoldering type.
A major complication of ATLL is the immunodeficiency of patients that leads to serious infections with
bacteria, fungi, protozoa, and viruses. Common infections include Pneumocystis carinii, aspergillosis, candidiasis, cytomegalovirus pneumonia, and Strongyloides
stercoralis [5661]. Opportunistic malignancies, such as
Kaposi sarcoma [62] and Epstein-Barr virus (EBV)associated lymphoma have been reported in patients
with ATL [63]. Although in the last 20 years considerable progress has been made regarding the biology of

Concise Review: Current Views in HTLV-I-Associated Adult T-Cell Leukemia/Lymphoma

HTLV-I, treatment of ATL patients remains unsatisfactory. ATL generally has a very poor prognosis and,
in leukemic or lymphomatous presentation, life expectancy does not exceed 1 year. High-dose radiotherapy
or chemotherapy regimens, by themselves or in combination, including those designed for the treatment of
aggressive non-Hodgkin lymphomas or acute lymphoblastic leukemia, are ineffective in ATL patients.
Although initial treatments often result in complete
remissions (40% CR), all patients relapse and die,
usually in less than a year. The poor prognosis of
ATL results from a combination of several factors.
Immune deficiency often results in opportunistic infections, and failure of hepatic functions prevents administration of intensive induction treatments. Other
negative factors include the intrinsic resistance of
ATL cells to apoptotic stimuli, often facilitated by
mutations in the p53 gene [64] and p16ink [65], overexpression of multidrug resistance [66], and the ATLderived factor (ADF), a thioredoxin analogue [67,68].
Many polychemotherapy clinical trials were carried out
in Japan between 1978 and 1983 [69], and although the
CR rate was usually higher in the lymphoma type as
compared to acute ATL, the long-term survival was
identical. Despite these obstacles recent encouraging
results were obtained by allogeneic bone marrow transplantation (alloBMT) [70], combinations of AZT and
a-IFN [7174], arsenic trioxide and a-IFN [75,76], alltrans-retinoic acid (ATRA) therapy [7779], and the
use of radiolabeled anti IL-2R (CD25) antibodies
[80,81]. Current therapeutic strategies for the treatment
of ATLL have recently been reviewed [82].
Molecular Pathogenesis

The low incidence and the long latency of HTLV-Iassociated ATLL suggest, in addition to viral infection, accumulations of genetic mutations are required
for cellular transformation in vivo. HTLV-I-mediated
T-cell transformation presumably arises from a multistep oncogenic process [83], in which the virus or
environmental factors induce chronic T-cell proliferation resulting in an accumulation of genetic defects
and the deregulated growth of infected cells. There is
a recognized discrepancy between the number of cells
carrying the provirus and the expression of viral
mRNA, even in the early stages of the disease [84].
Two possible models may explain this observation:
either cells are latently infected or cells expressing
viral antigens are rapidly eliminated by immune
responses. The higher viral load and polyclonal expansion of infected cells observed in TSP/HAM pathogenesis, suggest it may result from chronic virus
expression and a state of balance with the immune
system. However, to achieve the monoclonal expansion

235

that characterizes ATL [85], the virus has to establish a


latent reservoir that can be amplified by cellular replication. In fact, studies of clonal expansion of HTLVI-infected cells in HTLV-I carriers, demonstrate some
clones persist for over 7 years in the same individual
[86,87]. We have recently found that the virally encoded
p30 protein prevents nuclear export of the tax/rex
RNA and suppresses viral gene expression in vitro
[88]. Whether or not p30 may act as latency factor in
vivo remains to be investigated.
While it is not yet fully understood how HTLV-I
engenders ATLL, several lines of evidence have established that the viral oncoprotein Tax plays a central
role, at least in the early stages of the disease [89]. Tax
usurps signaling pathways, including NF-kappa B
[90], leading to the up-regulation of numerous cytokines and cytokine receptors. Remarkably, Tax has
been shown to up-regulate expression of IL-2 and
IL-2Ra chain [91,92] as well as IL-15 and IL-15Ra
chain [93,94], suggesting that an autocrine/paracrine
mechanism could be involved in proliferation of
ATL cells in early stages of infection [95]. Of note,
recent studies have found that the viral protein p12
stimulates production of IL-2 in activated T-cells [96],
interacts with the IL-2 receptor b chain, and stimulates the Jak/STAT5 pathway and T-cell proliferation
[97]. A more detailed role of p12 in HTLV-I pathogenesis has recently been reviewed [98]. A common
and striking feature of ATL cells in late stages of the
disease is the absence of detectable Tax expression,
suggesting that Tax expression may no longer be
required [99]. However, ATL cells appear to have
acquired a Tax phenotype: NF-kB and AP-1 are
constitutively activated [100,101], p53 is stabilized
and functionally impaired in the absence of mutations
[102], and expression of p21waf, survivin, and Bcl-xL
is increased [103106].
During the transformation process, Tax is also
involved in deregulation of cell growth [107] and apoptosis pathways [108,109], repression of the b-polymerase [110], the host DNA repair machinery [111], the
anaphase promoting complex [112], and inactivation
of the mitotic arrest defective (MAD1) protein [113].
These effects increase the occurrence and accumulation
of somatic mutations and predispose HTLV-I infected
cells to chromosome instability. As expected, reactivation of the human telomerase gene catalytic subunit
(hTERT) and increased telomerase activity is commonly found in HTLV-I infected cell lines in vitro
and in ex vivo ATL cells [114118]. Recent studies
demonstrate Tax, through its NF-kB inducing activity,
stimulates hTERT expression in HTLV-I-infected cells,
allowing maintenance of long telomeres and avoidance
of replicative senescence [118]. However, in the
presence of antigenic stimulation of T-lymphocytes

236

Concise Review: Nicot

bearing Tax, hTERT mRNA induction is diminished


[118,119]. hTERT is highly inducible in lymphocytes
following activation through CD3, possibly to maintain genetic stability of actively dividing cells. We propose the following model: in HTLV-I-infected T-cells,
Tax-mediated down-regulation of Lck, TCR, CD45,
and Syk/Zap-70 kinase expression results in attenuations of CD3 responses [120,121] and prevents the
full induction of hTERT expression following TCR
engagement. In turn, active cellular proliferation in
the presence of limiting amounts of telomerase may
result in a transient state of genetic instability. Once
the mitogenic effect has vanished, Tax-mediated activation of hTERT may stabilize and, thereafter, promote
the long-term expansion of potential tumor cells
that have acquired chromosomal abnormalities. Several cycles of transient active proliferation combined
with chromosomal instability may be required for a
clonal selection and the development of adult T-cell
leukemia. In support of such a model, a high frequency
of T-cell clonal expansion has been associated
with chronic antigenic stimulation in carriers of
S. stercoralis [122,123], and a higher frequency of leukemia has been reported in individuals carrying this
parasite [124126].
REFERENCES
1. Uchiyama T, Yodoi J, Sagawa K, Takatsuki K, Uchino H. Adult
T-cell leukemia: clinical and hematologic features of 16 cases.
Blood 1977;50:481492.
2. Mier JW, Gallo RC. Purification and some characteristics of
human T-cell growth factor from phytohemagglutinin-stimulated
lymphocyte-conditioned media. Proc Natl Acad Sci USA 1980;
77:61346138.
3. Poiesz BJ, Ruscetti FW, Gazdar AF, Bunn PA, Minna JD, Gallo
RC. Detection and isolation of type C retrovirus particles from
fresh and cultured lymphocytes of a patient with cutaneous T-cell
lymphoma. Proc Natl Acad Sci USA 1980;77:74157419.
4. Yoshida M, Miyoshi I, Hinuma Y. Isolation and characterization
of retrovirus from cell lines of human adult T-cell leukemia and its
implication in the disease. Proc Natl Acad Sci USA 1982;79:
20312035.
5. Aisenberg AC. Coherent view of non-Hodgkins lymphoma. J Clin
Oncol 1995;13:26562675.
6. Ohshima K, Suzumiya J, Kikuchi M. The World Health Organization classification of malignant lymphoma: incidence and clinical prognosis in HTLV-1-endemic area of Fukuoka. Pathol Int
2002;52:112.
7. Takahashi K, Takezaki T, Oki T, et al. Inhibitory effect of maternal
antibody on mother-to-child transmission of human T-lymphotropic
virus type I. The Mother-to-Child Transmission Study Group. Int J
Cancer 1991;49:673677.
8. Furnia A, Lal R, Maloney E, et al. Estimating the time of HTLV-I
infection following mother-to-child transmission in a breast-feeding population in Jamaica. J Med Virol 1999;59:541546.
9. Oki T, Yoshinaga M, Otsuka H, Miyata K, Sonoda S, Nagata Y.
A sero-epidemiological study on mother-to-child transmission of
HTLV-I in southern Kyushu, Japan. Asia Oceania J Obstet
Gynaecol 1992;18:371377.

10. Hino S, Sugiyama H, Doi H. [Mother-to-child transmission of


HTLV-I]. Nippon Rinsho 1986;44:22832288 (in Japanese).
11. Tajima K, Kamura S, Ito S, et al. Epidemiological features of
HTLV-I carriers and incidence of ATL in an ATL-endemic
island: a report of the community-based co-operative study in
Tsushima, Japan. Int J Cancer 1987;40:741746.
12. Take H, Umemoto M, Kusuhara K, Kuraya K. Transmission
routes of HTLV-I: an analysis of 66 families. Jpn J Cancer Res
1993;84:12651267.
13. Larson CJ, Taswell HF. Human T-cell leukemia virus type I
(HTLV-I) and blood transfusion. Mayo Clin Proc 1988;63:
869875.
14. Tobinai K, Shimoyama M. [Immunologic phenotype of malignant
lymphoma]. Nippon Rinsho 1992;50:12231228 (in Japanese).
15. Kamihira S, Sohda H, Atogami S, et al. Phenotypic diversity and
prognosis of adult T-cell leukemia. Leuk Res 1992;16:435441.
16. Sadamori N, Yamaguchi K, Ikeda S, et al. [Serum deoxythymidine kinase in adult T-cell leukemia and its related disorders].
Rinsho Ketsueki 1990;31:18121817 (in Japanese).
17. Sadamori N, Ichiba M, Mine M, et al. Clinical significance of
serum thymidine kinase in adult T-cell leukaemia and acute
myeloid leukaemia. Br J Haematol 1995;90:100105.
18. Kamihira S, Atogami S, Sohda H, Momita S, Yamada Y, Tomonaga M. Significance of soluble interleukin-2 receptor levels for
evaluation of the progression of adult T-cell leukemia. Cancer
1994;73:27532758.
19. Sadamori N, Mine M, Hakariya S, et al. Clinical significance of
beta 2-microglobulin in serum of adult T cell leukemia. Leukemia
1995;9:594597.
20. Fujiwara H, Arima N, Ohtsubo H, et al. Clinical significance of
serum neuron-specific enolase in patients with adult T-cell leukemia. Am J Hematol 2002;71:8084.
21. Shirono K, Hattori T, Takatsuki K. A new classification of
clinical stages of adult T-cell leukemia based on prognosis of
the disease. Leukemia 1994;8:18341837.
22. Itoyama T, Chaganti RS, Yamada Y, et al. Cytogenetic analysis
and clinical significance in adult T-cell leukemia/lymphoma: a
study of 50 cases from the human T-cell leukemia virus type-1
endemic area, Nagasaki. Blood 2001;97:36123620.
23. Smith SD, Morgan R, Link MP, McFall P, Hecht F. Cytogenetic
and immunophenotypic analysis of cell lines established from
patients with T cell leukemia/lymphoma. Blood 1986;67:650656.
24. Whang-Peng J, Bunn PA, Knutsen T, et al. Cytogenetic studies in
human T-cell lymphoma virus (HTLV)-positive leukemia-lymphoma in the United States. J Natl Cancer Inst 1985;74:357369.
25. Shimoyama M, Abe T, Miyamoto K, et al. Chromosome aberrations and clinical features of adult T cell leukemia-lymphoma not
associated with human T cell leukemia virus type I. Blood
1987;69:984989.
26. Hatta Y, Yamada Y, Tomonaga M, Miyoshi I, Said JW, Koeffler HP. Detailed deletion mapping of the long arm of chromosome 6 in adult T-cell leukemia. Blood 1999;93:613616.
27. Hall WW, Liu CR, Schneewind O, et al. Deleted HTLV-I provirus in blood and cutaneous lesions of patients with mycosis
fungoides. Science 1991;253:317320.
28. Khan ZM, Sebenik M, Zucker-Franklin D. Localization of
human T-cell lymphotropic virus-1 tax proviral sequences in
skin biopsies of patients with mycosis fungoides by in situ polymerase chain reaction. J Invest Dermatol 1996;106:667672.
29. Peterman A, Jerdan M, Staal S, et al. Evidence for HTLV-I
associated with mycosis fungoides and B-cell chronic lymphocytic leukemia. Arch Dermatol 1986;122:568571.
30. Zucker-Franklin D, Pancake BA, Friedman-Kien AE. Cutaneous disease resembling mycosis fungoides in HIV-infected
patients whose skin and blood cells also harbor proviral HTLV
type I. AIDS Res Hum Retroviruses 1994;10:11731177.

Concise Review: Current Views in HTLV-I-Associated Adult T-Cell Leukemia/Lymphoma


31. Zucker-Franklin D, Kosann MK, Pancake BA, Ramsay DL,
Soter NA. Hypopigmented mycosis fungoides associated with
human T cell lymphotropic virus type I tax in a pediatric patient.
Pediatrics 1999;103:10391045.
32. Zucker-Franklin D. The role of human T cell lymphotropic virus
type I tax in the development of cutaneous T cell lymphoma. Ann
NY Acad Sci 2001;941:8696.
33. Pancake BA, Zucker-Franklin D, Coutavas EE. The cutaneous T
cell lymphoma, mycosis fungoides, is a human T cell lymphotropic virus-associated disease. A study of 50 patients. J Clin Invest
1995;95:547554.
34. Zucker-Franklin D, Coutavas EE, Rush MG, Zouzias DC.
Detection of human T-lymphotropic virus-like particles in cultures of peripheral blood lymphocytes from patients with mycosis
fungoides. Proc Natl Acad Sci USA 1991;88:76307634.
35. Barbosa HS, Bittencourt AL, Barreto de Araujo I, et al. Adult Tcell leukemia/lymphoma in northeastern Brazil: a clinical, histopathologic, and molecular study. J Acquir Immune Defic Syndr
1999;21:6571.
36. Kikuchi A, Nishikawa T, Yamaguchi K. Absence of human
T-cell lymphotropic virus type I in cutaneous T-cell lymphoma.
N Engl J Med 1997;336:296297.
37. Lisby G, Reitz MS Jr, Vejlsgaard GL. No detection of HTLV-I
DNA in punch skin biopsies from patients with cutaneous T-cell
lymphoma by the polymerase chain reaction. J Invest Dermatol
1992;98:417420.
38. Bazarbachi A, Soriano V, Pawson R, et al. Mycosis fungoides
and Sezary syndrome are not associated with HTLV-I infection:
an international study. Br J Haematol 1997;98:927933.
39. Capesius C, Saal F, Maero E, et al. No evidence for HTLV-I infection in 24 cases of French and Portuguese mycosis fungoides and
Sezary syndrome (as seen in France). Leukemia 1991;5:416419.
40. Lapis P, Freeman J, Bitter MA, Golitz LE. Absence of HTLV-I
DNA sequences in cutaneous T-cell lymphoma/mycosis fungoides. Acta Morphol Hung 1992;40:249255.
41. Dosaka N, Tanaka T, Miyachi Y, Imamura S, Kakizuka A.
Examination of HTLV-I integration in the skin lesions of various
types of adult T-cell leukemia (ATL): independence of cutaneous-type ATL confirmed by Southern blot analysis. J Invest
Dermatol 1991;96:196200.
42. Gessain A, Moulonguet I, Flageul B, et al. Cutaneous type of adult
T cell leukemia/lymphoma in a French West Indian woman. Clonal
rearrangement of T-cell receptor beta and gamma genes and monoclonal integration of HTLV-I proviral DNA in the skin infiltrate.
J Am Acad Dermatol 1990;23:9941000.
43. Hamada T, Setoyama M, Katahira Y, et al. Differences in HTLV-I
integration patterns between skin lesions and peripheral blood
lymphocytes of HTLV-I seropositive patients with cutaneous lymphoproliferative disorders. J Dermatol Sci 1992;4:7682.
44. Whittaker SJ, Ng YL, Rustin M, Levene G, McGibbon DH,
Smith NP. HTLV-1-associated cutaneous disease: a clinicopathological and molecular study of patients from the U.K. Br J
Dermatol 1993;128:483492.
45. Arai M, Ohashi T, Tsukahara T, et al. Human T-cell leukemia
virus type 1 Tax protein induces the expression of lymphocyte
chemoattractant SDF-1/PBSF. Virology 1998;241:298303.
46. Yoshie O, Fujisawa R, Nakayama T, et al. Frequent expression
of CCR4 in adult T-cell leukemia and human T-cell leukemia
virus type 1-transformed T cells. Blood 2002;99:15051511.
47. Yamaguchi T, Ohshima K, Tsuchiya T, et al. The comparison of
expression of cutaneous lymphocyte-associated antigen (CLA), and
Th1- and Th2-associated antigens in mycosis fungoides and cutaneous lesions of adult T-cell leukemia/lymphoma. Eur J Dermatol
2003;13:553559.
48. Berg EL, Yoshino T, Rott LS, et al. The cutaneous lymphocyte
antigen is a skin lymphocyte homing receptor for the vascular

49.

50.

51.

52.

53.

54.

55.

56.

57.

58.

59.

60.

61.

62.

63.

64.

65.

66.

67.

237

lectin endothelial cell-leukocyte adhesion molecule 1. J Exp Med


1991;174:14611466.
Bertini R, Howard OM, Dong HF, et al. Thioredoxin, a redox
enzyme released in infection and inflammation, is a unique chemoattractant for neutrophils, monocytes, and T cells. J Exp Med
1999;189:17831789.
Barmak K, Harhaj E, Grant C, Alefantis T, Wigdahl B. Human
T cell leukemia virus type I-induced disease: pathways to cancer
and neurodegeneration. Virology 2003;308:112.
Grant C, Barmak K, Alefantis T, Yao J, Jacobson S, Wigdahl B.
Human T cell leukemia virus type I and neurologic disease:
events in bone marrow, peripheral blood, and central nervous
system during normal immune surveillance and neuroinflammation. J Cell Physiol 2002;190:133159.
Kiwaki T, Umehara F, Arimura Y, et al. The clinical and pathological features of peripheral neuropathy accompanied with
HTLV-I associated myelopathy. J Neurol Sci 2003;206:1721.
Shimoyama M. Diagnostic criteria and classification of clinical
subtypes of adult T-cell leukaemia-lymphoma. A report from the
Lymphoma Study Group (1984-87). Br J Haematol 1991;79:
428437.
Honda S, Yamaguchi K, Miyake Y, et al. Production of parathyroid hormone-related protein in adult T-cell leukemia cells.
Jpn J Cancer Res 1988;79:12641268.
Reichel H, Koeffler HP, Norman AW. 25-Hydroxyvitamin D3
metabolism by human T-lymphotropic virus-transformed lymphocytes. J Clin Endocrinol Metab 1987;65:519526.
Newton RC, Limpuangthip P, Greenberg S, Gam A, Neva FA.
Strongyloides stercoralis hyperinfection in a carrier of HTLV-I
virus with evidence of selective immunosuppression. Am J Med
1992;92:202208.
Grossman ME, Pappert AS, Garzon MC, Silvers DN. Invasive
Trichophyton rubrum infection in the immunocompromised host:
report of three cases. J Am Acad Dermatol 1995;33:315318.
Pagliuca A, Layton DM, Allen S, Mufti GJ. Hyperinfection with
Strongyloides after treatment for adult T cell leukaemialymphoma
in an African immigrant. Br Med J 1988;297:14561457.
Yamamoto N, Miyara T, Kawakami K, et al. [A case of disseminated aspergillosis with smoldering adult T-cell leukemia].
Kansenshogaku Zasshi 2002;76:460465 (in Japanese).
Obata S, Matsuzaki H, Nishimura H, Kawakita M, Takatsuki K.
Gastroduodenal complications in patients with adult T-cell leukemia. Jpn J Clin Oncol 1988;18:335342.
Roudier M, Lamaury I, Strobel M. Human T cell leukemia/
lymphoma virus type I (HTLV-I) and Pneumocystis carinii associated with T cell proliferation and haemophagocytic syndrome.
Leukemia 1997;11:453454.
Greenberg SJ, Jaffe ES, Ehrlich GD, Korman NJ, Poiesz BJ,
Waldmann TA. Kaposis sarcoma in human T-cell leukemia virus
type I-associated adult T-cell leukemia. Blood 1990;76:971976.
Tobinai K, Ohtsu T, Hayashi M, et al. Epstein-Barr virus (EBV)
genome carrying monoclonal B-cell lymphoma in a patient with
adult T-cell leukemia-lymphoma. Leuk Res 1991;15:837846.
Nagai H, Kinoshita T, Imamura J, et al. Genetic alteration of
p53 in some patients with adult T-cell leukemia. Jpn J Cancer Res
1991;82:14211427.
Hatta Y, Hirama T, Miller CW, Yamada Y, Tomonaga M,
Koeffler HP. Homozygous deletions of the p15 (MTS2) and
p16 (CDKN2/MTS1) genes in adult T-cell leukemia. Blood
1995;85:26992704.
Lau A, Nightingale S, Taylor GP, Gant TW, Cann AJ. Enhanced
MDR1 gene expression in human T-cell leukemia virus-I-infected
patients offers new prospects for therapy. Blood 1998;91:
24672474.
Yodoi J, Tursz T. ADF, a growth-promoting factor derived from
adult T cell leukemia and homologous to thioredoxin: involvement

238

68.

69.

70.

71.

72.

73.

74.

75.

76.

77.

78.

79.

80.

81.

82.

83.

84.

Concise Review: Nicot


in lymphocyte immortalization by HTLV-I and EBV. Adv Cancer
Res 1991;57:381411.
Tagaya Y, Maeda Y, Mitsui A, et al. ATL-derived factor (ADF),
an IL-2 receptor/Tac inducer homologous to thioredoxin; possible involvement of dithiol-reduction in the IL-2 receptor induction. EMBO J 1989;8:757764.
Tsukasaki K, Tobinai K, Shimoyama M, et al. Deoxycoformycincontaining combination chemotherapy for adult T-cell leukemialymphoma: Japan Clinical Oncology Group Study (JCOG9109).
Int J Hematol 2003;77:164170.
Utsunomiya A, Miyazaki Y, Takatsuka Y, et al. Improved outcome of adult T cell leukemia/lymphoma with allogeneic hematopoietic stem cell transplantation. Bone Marrow Transplant 2001;
27:1520.
Bazarbachi A, Hermine O. Treatment with a combination of
zidovudine and alpha-interferon in naive and pretreated adult
T-cell leukemia/lymphoma patients. J Acquir Immune Defic
Syndr Hum Retrovirol 1996;13(Suppl 1):S186S190.
Hermine O, Bouscary D, Gessain A, et al. Brief report: treatment
of adult T-cell leukemia-lymphoma with zidovudine and interferon alfa. N Engl J Med 1995;332:17491751.
Gill PS, Harrington W Jr, Kaplan MH, et al. Treatment of adult
T-cell leukemia-lymphoma with a combination of interferon alfa
and zidovudine. N Engl J Med 1995;332:17441748.
Tobinai K, Kobayashi Y, Shimoyama M. Interferon alfa and
zidovudine in adult T-cell leukemia-lymphoma. Lymphoma
Study Group of the Japan Clinical Oncology Group. N Engl J
Med 1995;333:1285.
Bazarbachi A, El Sabban ME, Nasr R, et al. Arsenic trioxide and
interferon-alpha synergize to induce cell cycle arrest and apoptosis in human T-cell lymphotropic virus type I-transformed cells.
Blood 1999;93:278283.
Mahieux R, Pise-Masison C, Gessain A, et al. Arsenic trioxide
induces apoptosis in human T-cell leukemia virus type 1- and
type 2-infected cells by a caspase-3-dependent mechanism involving Bcl-2 cleavage. Blood 2001;98:37623769.
Maeda Y, Miyatake J, Sono H, et al. 13-cis-Retinoic acid inhibits
growth of adult T cell leukemia cells and causes apoptosis; possible
new indication for retinoid therapy. Intern Med 1996;35:180184.
Miyatake J, Maeda Y, Nawata H, et al. Thiol compounds rescue
growth inhibition by retinoic acid on HTLV-I (+) T lymphocytes; possible mechanism of retinoic-acid-induced growth inhibition of adult T-cell leukemia cells. Hematopathol Mol Hematol
1998;11:8999.
Nawata H, Maeda Y, Sumimoto Y, Miyatake J, Kanamaru A. A
mechanism of apoptosis induced by all-trans-retinoic acid on adult
T-cell leukemia cells: a possible involvement of the Tax/NF-kappaB signaling pathway. Leuk Res 2001;25:323331.
Waldmann TA. T-cell receptors for cytokines: targets for immunotherapy of leukemia/lymphoma. Ann Oncol 2000;11(Suppl 1):
101106.
Waldmann TA, White JD, Goldman CK, et al. The interleukin-2
receptor: a target for monoclonal antibody treatment of human
T-cell lymphotrophic virus I-induced adult T-cell leukemia.
Blood 1993;82:17011712.
Bazarbachi A, Hermine O. Treatment of adult T-cell leukaemia/
lymphoma: current strategy and future perspectives. Virus Res
2001;78:7992.
Okamoto T, Ohno Y, Tsugane S, et al. Multi-step carcinogenesis
model for adult T-cell leukemia. Jpn J Cancer Res 1989;80:
191195.
Gessain A, Louie A, Gout O, Gallo RC, Franchini G. Human Tcell leukemia-lymphoma virus type I (HTLV-I) expression in
fresh peripheral blood mononuclear cells from patients with tropical spastic paraparesis/HTLV-I-associated myelopathy. J Virol
1991;65:16281633.

85. Wattel E, Cavrois M, Gessain A, Wain-Hobson S. Clonal expansion of infected cells: a way of life for HTLV-I. J Acquir Immune
Defic Syndr Hum Retrovirol 1996;13(Suppl 1):S92S99.
86. Cavrois M, Leclercq I, Gout O, Gessain A, Wain-Hobson S,
Wattel E. Persistent oligoclonal expansion of human T-cell
leukemia virus type 1-infected circulating cells in patients with
tropical spastic paraparesis/HTLV-1 associated myelopathy.
Oncogene 1998;17:7782.
87. Etoh K, Tamiya S, Yamaguchi K, et al. Persistent clonal proliferation of human T-lymphotropic virus type I-infected cells in
vivo. Cancer Res 1997;57:48624867.
88. Nicot C, Dundr M, Johnson JM, et al. HTLV-1-encoded p30II is
a post-transcriptional negative regulator of viral replication. Nat
Med 2004;10:197201.
89. Franchini G, Nicot C, Johnson JM. Seizing of T cells by human
T-cell leukemia/lymphoma virus type 1. Adv Cancer Res 2003;89:
69132.
90. Sun SC, Harhaj EW, Xiao G, Good L. Activation of I-kappaB
kinase by the HTLV type 1 Tax protein: mechanistic insights into
the adaptor function of IKKgamma. AIDS Res Hum Retroviruses 2000;16:15911596.
91. Ruben S, Poteat H, Tan TH, et al. Cellular transcription factors
and regulation of IL-2 receptor gene expression by HTLV-I tax
gene product. Science 1988;241:8992.
92. Good L, Maggirwar SB, Sun SC. Activation of the IL-2 gene
promoter by HTLV-I tax involves induction of NF-AT complexes bound to the CD28-responsive element. EMBO J 1996;
15:37443750.
93. Mariner JM, Lantz V, Waldmann TA, Azimi N. Human T cell
lymphotropic virus type I Tax activates IL-15R alpha gene expression through an NF-kappa B site. J Immunol 2001;166: 26022609.
94. Waldmann TA. The promiscuous IL-2/IL-15 receptor: a target
for immunotherapy of HTLV-I-associated disorders. J Acquir
Immune Defic Syndr Hum Retrovirol 1996;13(Suppl 1):
S179S185.
95. Russell SJ. Interleukin-2 and T cell malignancies: an autocrine
loop with a twist. Leukemia 1989;3:755757.
96. Ding W, Kim SJ, Nair AM, et al. Human T-cell lymphotropic
virus type 1 p12(I) enhances interleukin-2 production during
T-cell activation. J Virol 2003;77:1102711039.
97. Nicot C, Mulloy JC, Ferrari MG, et al. HTLV-1 p12(I) protein
enhances STAT5 activation and decreases the interleukin-2
requirement for proliferation of primary human peripheral
blood mononuclear cells. Blood 2001;98:823829.
98. Albrecht B, Lairmore MD. Critical role of human T-lymphotropic virus type 1 accessory proteins in viral replication and pathogenesis. Microbiol Mol Biol Rev 2002;66:396406.
99. Bangham CR. Human T-lymphotropic virus type 1 (HTLV-1):
persistence and immune control. Int J Hematol 2003;78:297303.
100. Mori N, Fujii M, Iwai K, et al. Constitutive activation of transcription factor AP-1 in primary adult T-cell leukemia cells.
Blood 2000;95:39153921.
101. Mori N, Fujii M, Ikeda S, et al. Constitutive activation of NFkappaB in primary adult T-cell leukemia cells. Blood 1999;93:
23602368.
102. Takemoto S, Trovato R, Cereseto A, et al. p53 stabilization and
functional impairment in the absence of genetic mutation or the
alteration of the p14(ARF)-MDM2 loop in ex vivo and cultured
adult T-cell leukemia/lymphoma cells. Blood 2000;95:39393944.
103. Nicot C, Mahieux R, Takemoto S, Franchini G. Bcl-X(L) is upregulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL
samples. Blood 2000;96:275281.
104. Kamihira S, Yamada Y, Hirakata Y, et al. Aberrant expression
of caspase cascade regulatory genes in adult T-cell leukaemia:
survivin is an important determinant for prognosis. Br J Haematol 2001;114:6369.

Concise Review: Current Views in HTLV-I-Associated Adult T-Cell Leukemia/Lymphoma


105. Mori N, Yamada Y, Hata T, et al. Expression of survivin in
HTLV-I-infected T-cell lines and primary ATL cells. Biochem
Biophys Res Commun 2001;282:11101113.
106. de La Fuente C, Santiago F, Chong SY, et al. Overexpression of
p21(waf1) in human T-cell lymphotropic virus type 1-infected
cells and its association with cyclin A/cdk2. J Virol 2000;74:
72707283.
107. Gatza ML, Watt JC, Marriott SJ. Cellular transformation by the
HTLV-I Tax protein, a jack-of-all-trades. Oncogene 2003;22:
51415149.
108. Nicot C, Harrod R. Distinct p300-responsive mechanisms promote caspase-dependent apoptosis by human T-cell lymphotropic virus type 1 Tax protein. Mol Cell Biol 2000;20:85808589.
109. de La Fuente C, Wang L, Wang D, et al. Paradoxical effects of a
stress signal on pro- and anti-apoptotic machinery in HTLV-1
Tax expressing cells. Mol Cell Biochem 2003;245:99113.
110. Jeang KT, Widen SG, Semmes OJ, Wilson SH. HTLV-I transactivator protein, tax, is a trans-repressor of the human betapolymerase gene. Science 1990;247:10821084.
111. Lemoine FJ, Kao SY, Marriott SJ. Suppression of DNA repair
by HTLV type 1 Tax correlates with Tax trans-activation of
proliferating cell nuclear antigen gene expression. AIDS Res
Hum Retroviruses 2000;16:16231627.
112. Liu B, Liang MH, Kuo YL, et al. Human T-lymphotropic virus
type 1 oncoprotein tax promotes unscheduled degradation of
Pds1p/securin and Clb2p/cyclin B1 and causes chromosomal
instability. Mol Cell Biol 2003;23:52695281.
113. Jin DY, Spencer F, Jeang KT. Human T cell leukemia virus type
1 oncoprotein Tax targets the human mitotic checkpoint protein
MAD1. Cell 1998;93:8191.
114. Franzese O, Balestrieri E, Comandini A, Forte G, Macchi B,
Bonmassar E. Telomerase activity of human peripheral blood
mononuclear cells in the course of HTLV type 1 infection in
vitro. AIDS Res Hum Retroviruses 2002;18:249251.
115. Re MC, Monari P, Gibellini D, et al. Human T cell leukemia
virus type II increases telomerase activity in uninfected CD34+
hematopoietic progenitor cells. J Hematother Stem Cell Res
2000;9:481487.

239

116. Tsumuki H, Nakazawa M, Hasunuma T, et al. Infection of


synoviocytes with HTLV-I induces telomerase activity. Rheumatol Int 2001;20:175179.
117. Uchida N, Otsuka T, Arima F, et al. Correlation of telomerase
activity with development and progression of adult T-cell leukemia. Leuk Res 1999;23:311316.
118. Sinha-Datta U, Horikawa I, Michishita E, et al. Transcriptional
activation of hTERT through the NF-kB pathway in HTLV-I
transformed cells. Blood 2004;104(8):25232531.
119. Gabet AS, Mortreux F, Charneau P, et al. Inactivation of
hTERT transcription by Tax. Oncogene 2003;22:37343741.
120. Weil R, Levraud JP, Dodon MD, et al. Altered expression of
tyrosine kinases of the Src and Syk families in human T-cell
leukemia virus type 1-infected T-cell lines. J Virol 1999;73:
37093717.
121. Lemasson I, Robert-Hebmann V, Hamaia S, Duc DM, Gazzolo L,
Devaux C. Transrepression of lck gene expression by human
T-cell leukemia virus type 1-encoded p40tax. J Virol 1997;71:
19751983.
122. Agape P, Copin MC, Cavrois M, et al. Implication of HTLV-I
infection, strongyloidiasis, and P53 overexpression in the development, response to treatment, and evolution of non-Hodgkins
lymphomas in an endemic area (Martinique, French West
Indies). J Acquir Immune Defic Syndr Hum Retrovirol 1999;20:
394402.
123. Gabet AS, Mortreux F, Talarmin A, et al. High circulating
proviral load with oligoclonal expansion of HTLV-1 bearing T
cells in HTLV-1 carriers with strongyloidiasis. Oncogene 2000;
19:49544960.
124. Nakada K, Yamaguchi K, Furugen S, et al. Monoclonal integration of HTLV-I proviral DNA in patients with strongyloidiasis.
Int J Cancer 1987;40:145148.
125. Yamaguchi K, Matutes E, Catovsky D, Galton DA, Nakada K,
Takatsuki K. Strongyloides stercoralis as candidate co-factor for
HTLV-I-induced leukaemogenesis. Lancet 1987;2:9495.
126. Plumelle Y, Gonin C, Edouard A, et al. Effect of Strongyloides
stercoralis infection and eosinophilia on age at onset and prognosis of adult T-cell leukemia. Am J Clin Pathol 1997;107:8187.

You might also like